Вы находитесь на странице: 1из 8

Colloids and Surfaces A: Physicochem. Eng.

Aspects 414 (2012) 220–227

Contents lists available at SciVerse ScienceDirect

Colloids and Surfaces A: Physicochemical and


Engineering Aspects
journal homepage: www.elsevier.com/locate/colsurfa

Removal of uranium(VI) from aqueous solutions by magnetic Mg–Al layered


double hydroxide intercalated with citrate: Kinetic and thermodynamic
investigation
Xiaofei Zhang a , Lanyang Ji b , Jun Wang a,c,∗ , Rumin Li a , Qi Liu a , Milin Zhang a , Lianhe Liu c
a
Key Laboratory of Superlight Material and Surface Technology, Ministry of Education, Harbin Engineering University, Harbin 150001, China
b
Hygienic Inspection and Monitor Station for Cereals and Oil of Heilongjiang Province, 150001, China
c
Institute of Advanced Marine Materials, Harbin Engineering University, 150001, China

h i g h l i g h t s g r a p h i c a l a b s t r a c t

 We prepared magnetic citrate·Mg–Al


LDH via ion-exchange technique.
 It can effectively removal ura-
nium(VI) from aqueous solutions.
 Adsorption product is easily sepa-
rated by the external magnetic field.
 The maximum adsorption capacity
toward uranium(VI) is 180.00 mg g−1
at 25 ◦ C.

a r t i c l e i n f o a b s t r a c t

Article history: Magnetic Mg–Al layered double hydroxide was modified with citrate acid using a ion-exchange
Received 1 June 2012 technique. It showed higher adsorption capacity for the uranium(VI) ions compared with magnetic
Received in revised form 2 August 2012 Mg–Al layered double hydroxide without intercalating with citrate. This increasing capacity was
Accepted 5 August 2012
mainly attributable to the formation of the citrate–metal complexes in the interlayer of the magnetic
Available online 27 August 2012
citrate·Mg–Al layered double hydroxide. Batch experiments were conducted to study the effects of pH,
adsorbent dose, shaking time, and temperature on uranium sorption efficiency. The results reveal that
Keywords:
the maximum adsorption capacity is 180.00 mg g−1 when the initial uranium(VI) concentration is kept
Uranium
Magnetic Mg–Al layered double hydroxide
as 200 mg L−1 at 25 ◦ C, displaying a high efficiency for the removal of uranium(VI) from aqueous solution.
Citrate The sorption follows Freundlich model and pseudo-second-order kinetics. The thermodynamic param-
Adsorption eters such as H◦ , S◦ and G◦ show that the process is endothermic and spontaneous. In addition,
Kinetics the composite can be easily separated from the solution by a magnet after the adsorption process. This
Thermodynamics work provides an efficient, fast and convenient approach for the removal of uranium(VI) from aqueous
solutions.
© 2012 Elsevier B.V. All rights reserved.

1. Introduction biological toxicity. Therefore, it is very important to choose


a suitable and effective method to removal of uranium from
Uranium is one of the most dangerous heavy metals in the water. So far, several methods, such as chemical precipitation,
environment because of its long half-life, high radioactivity and membrane separation, microbial removal, solvent extraction,
electrodialysis, and adsorption [1–8], have been extensively
applied for the removal of uranium(VI) from aqueous solutions.
∗ Corresponding author at: Key Laboratory of Superlight Material and Surface
Among these methods, adsorption appears to be one of the
Technology, Ministry of Education, Harbin Engineering University, Harbin 150001,
most effective methods owning to its cost-effective, versatile
China. Tel.: +86 451 8253 3026; fax: +86 451 8253 3026. and simple features to operate for removing trace levels of ions
E-mail address: zhqw1888@sohu.com (J. Wang). [9,10].

0927-7757/$ – see front matter © 2012 Elsevier B.V. All rights reserved.
http://dx.doi.org/10.1016/j.colsurfa.2012.08.031
X. Zhang et al. / Colloids and Surfaces A: Physicochem. Eng. Aspects 414 (2012) 220–227 221

Different types of adsorbents have been developed and tested


for removal of uranium(VI) from aqueous media [11–13]. Layered Fe3O4
double hydroxides (LDHs), because of their structure and high
anionic exchange capacity, have received considerable attention (b)

Intensity (a. u.)

311
in recent years. They have many potential applications such as sor-
bents, catalyst and catalyst supports [14–16]. Martinez-Gallegos
et al. [17] have used thermally treated hydrotalcite to immobilize
the CrO4 2− from aqueous solutions and understood the structural
effects of ␥-irradiation. Kulyukhin et al. [18] have synthesized of

440
220

400

511
Mg, Al and Mg, Nd layered double hydroxides for removal of ura-

422
111

622
nium from aqueous solutions. Just like LDH can be used as sorbents
and sorbents supports, it will be also used as catalyst and catalyst (a)
supports. Vijaikumar et al. [19] have immobilized L-proline onto the
inter layers of hydrotalcite clay and studied its catalytic application
in the asymmetric Michael addition reaction. 10 20 30 40 50 60 70 80
LDHs has an anion-exchange capability and is represented by
the formula [M1−x 2+ Mx 3+ (OH)2 ](An− )x/n ·mH2 O, where M2+ could be
2 degree
Mg2+ , Ni2+ , Zn2+ , etc.; M3+ could be Al3+ , Fe3+ , etc.; An− represents
Fig. 1. XRD patterns of magnetic Mg–Al LDH (a) and magnetic citrate·Mg–Al LDH
interlayer anions, such as NO3 − , SO4 2− , and CO3 2− , and x typically (b).
ranges from 0.20 to 0.33 [20]. Since the interlayer anions are easily
exchangeable, the anions in the interlayer can be exchanged by
2.2 g citric acid added to a suspension of magnetic Mg–Al LDH
other anions so that it make LDHs useful for many applications,
(1.2 g) in 25 mL distilled water at 50 ◦ C. This solution was added
such as the sorption of many hazardous inorganic and organic ions
dropwise to 25 mL (1 mol L−1 ) NaOH solution, followed by refluxing
from aqueous solution.
for 3 h. The magnetic citrate·Mg–Al LDH was recovered by filtration,
Recently, some authors successfully removed metal cations
washed, and dried at 70 ◦ C for 16 h.
from aqueous solutions by LDHs intercalated with ethylene
diamine tetraacetic acid (EDTA) [21,22] and nitrilotriacetate (NTA)
[23]. 2.2. Adsorption experiments
In this paper, we prepared magnetic Mg–Al LDH intercalated
with citrate using a ion-exchange technique (labeled as magnetic The adsorption of uranium(VI) from the dilute aqueous solu-
citrate·Mg–Al LDH). Citric acid exists abundantly in nature, and tion was operated as the following procedure: 0.05 g magnetic
it is found to be effective for the leaching of heavy metals due citrate·Mg–Al LDH was added to 50 mL of UO2 (NO3 )2 ·6H2 O solu-
to the formation of citrate–metal complexes between metals and tion. This mixture was shaken on a constant temperature oscillator
acids [24]. This work aims at preparing an adsorbent to effectively at 120 strokes min−1 . The solution pH was adjusted with 0.5 mol L−1
remove uranium(VI) from aqueous solutions. The effects of differ- HNO3 or NaOH solution. At the end of the adsorption period, the
ent parameters on uranium(VI) adsorption, such as solution pH solution was separated from the solids by magnetic separation.
value, adsorbent dose, contact time and the temperature have been Then the initial and the equilibrium concentration of tested ion in
studied. The adsorption kinetics, isotherms, and thermodynamics supernatant were determined. The adsorption capacity Qe (mg g−1 )
have also been investigated. was calculated according to Eq. (1):
(C0 − Ce )V
Qe = (1)
m
2. Experimental
where C0 (mg L−1 ) is the uranium(VI) ion concentration in the
2.1. Preparation of samples initial solution, Ce (mg L−1 ) is the equilibrium concentration of ura-
nium(VI) ion in the supernatant, V (L) is the volume of the testing
The magnetic particles were prepared according to the pre- solution and m is the weight of sorbent (g).
vious report [25]. 5.20 g of FeCl3 ·6H2 O, 2.7969 g of FeSO4 ·7H2 O
and 0.85 mL of concentrated HCl were dissolved in 200 mL of 2.3. Characterization
deoxygenated water. Subsequently, the solution was slowly added
dropwise into the previously prepared 250 mL of NaOH solu- X-ray diffraction (XRD) analysis was performed on a Rigaku
tion (0.75 mol L−1 ). Throughout the reaction process, the solution D/max-IIIB diffractometer with Cu K␣ irradiation ( = 1.54178 Å).
was stirred vigorously at 85 ◦ C under N2 protection. The reac- The X-ray source was operated at 40 kV and 150 mA. Fourier-
tion continued for another 25 min, and the mixture was cooled to transform infrared (FT-IR) spectrum was recorded with an AVATAR
room temperature. The black solid was washed several times with 360 FT-IR spectrophotometer using a standard KBr pellets. Effluent
ethanol and dried in vacuum at 60 ◦ C for 12 h. was analyzed using a Trace Uranium Analyzer.
Magnetic Mg–Al LDH was synthesized by a modified copre-
cipitation method as described elsewhere [26,27]. 0.10 mol L−1 3. Results and discussions
Mg(NO3 )2 and 0.05 mol L−1 Al(NO3 )3 were dissolved in deionized
water to prepare a mixed solution. Then, 0.4 g Fe3 O4 was introduced 3.1. Characterization of samples
in this mixed solution and followed by ultrasonic for 10 min. The
solution pH was adjusted to 10.5 by addition of 0.5 mol L−1 NaOH XRD patterns of magnetic Mg–Al LDH and magnetic
and 0.015 mol L−1 Na2 CO3 mixed solution. After the addition of the citrate·Mg–Al LDH are shown in Fig. 1. From the XRD pattern
solution, the mixed solution was kept stirring at 30 ◦ C for 1 h at a of the magnetic Mg–Al LDH, except the diffraction peaks of the
constant pH of 10.5. The magnetic Mg–Al LDH was recovered by fil- LDH, the peaks marked with  of (1 1 1), (2 2 0), (3 1 1), (4 0 0),
tering the resultant suspension and washed with deionized water, (4 2 2), (5 1 1), (4 4 0), and (6 2 2) planes can be indexed to cubic
and then dried in a vacuum oven at 50 ◦ C for 36 h. Fe3 O4 (JCPDS: 65-3107). Comparison of the XRD patterns finds that
222 X. Zhang et al. / Colloids and Surfaces A: Physicochem. Eng. Aspects 414 (2012) 220–227

100 200 b b
b
b a
a a
80 160 a
2895
Transmittance

Qe (mg/g)
60 120

40
80
3462 1579 623
20
1392
40
0
0
4000 3500 3000 2500 2000 1500 1000 500 25oC 35oC 45oC 55oC
wavenumber (cm-1) Temperature

Fig. 2. FT-IR spectra of magnetic citrate·Mg–Al LDH. Fig. 4. Adsorption capacity of uranium by magnetic Mg–Al LDH (a) and magnetic
citrate·Mg–Al LDH (b). Initial uranium concentration 200 mg L−1 ; pH 6.0; tempera-
ture 25–55 ◦ C; amount of adsorbent 0.05 g.

the XRD peaks for (b) are weakened and broadened at 2 values C H stretching vibrations. These results show that the citrate is
compared to those for (a). However, the 2 values corresponding present in the magnetic citrate·Mg–Al LDH.
to all of the XRD peaks for (b) are close to those for (a), suggesting The magnetic citrate·Mg–Al LDH was dispersed in water by vig-
that the magnetic citrate·Mg–Al LDH has the hydrotalcite lattice orous shaking, resulting in a brown suspension (Fig. 3A), and it
structure. For magnetic Mg–Al LDH, the observed basal spacing, could be easily removed from the aqueous solution with an addition
d0 0 3 , is 7.7 Å. For magnetic citrate·Mg–Al LDH, intercalation of of a permanent magnet, indicating the sensitive magnetic response
the chelating agents in CO3 ·Mg–Al LDH increased d0 0 3 from 7.7 Å (Fig. 3B). It can provide an easy and efficient way of separating the
to 11.6 Å, the broadened XRD peaks corresponding to the basal adsorbent from a suspension system during the sorption experi-
spacing are located at lower 2 values than those of CO3 ·Mg–Al ments.
LDH, also indicative of a larger interlayer spacing. These results
suggest that citrate, which is larger than CO3 2− , was intercalated 3.2. Uranium(VI)-sorption experiments
into the interlayer of the Mg–Al LDH.
Fig. 2 gives the FT-IR spectra of magnetic citrate·Mg–Al LDH. By using the magnetic Mg–Al LDH and the magnetic
The peaks at 623 cm−1 is assigned to Fe O bond vibration of citrate·Mg–Al LDH, the adsorption experiments were carried out for
Fe3 O4 . The intense and broad peak at 3462 cm−1 is ascribed to uranium(VI) at 25, 35, 45 and 55 ◦ C, respectively. As shown in Fig. 4,
the stretching vibration of hydroxyl groups of LDH layers and it can be seen that the adsorption ability of magnetic citrate·Mg–Al
interlayer water molecules. The absorption bands at 1392 and LDH particles for uranium(VI) ions is higher than magnetic Mg–Al
1579 cm−1 are assigned to the characteristic symmetric and LDH. By comparing the results for magnetic citrate·Mg–Al LDH with
asymmetric stretching vibrations of COO− , respectively [27]. In those for magnetic Mg–Al LDH, we can attribute the increasing
addition, the characteristic peak at 2895 cm−1 , attributes to the capacity of uranium(VI) by the magnetic citrate·Mg–Al LDH to the

Fig. 3. Digital photograph of the aqueous solution (A) with dispersed magnetic magnetic citrate·Mg–Al LDH composite particles and (B) after magnetic separation using an
external magnetic field.
X. Zhang et al. / Colloids and Surfaces A: Physicochem. Eng. Aspects 414 (2012) 220–227 223

200 190

180 180

160 170

Qe (mg/g)
Qe (mg/g)

140 160

120 150

100 140

80 130
2 4 6 8 10 12 0.00 0.02 0.04 0.06 0.08 0.10 0.12 0.14 0.16
pH Mass of sorbent (g)
Fig. 5. Effect of pH value on adsorption property of magnetic citrate·Mg–Al LDH. Fig. 6. The effect of adsorbent dose on the uptake of uranium(VI) by magnetic
Initial uranium concentration 200 mg L−1 ; pH 2.0–12.0; temperature 25 ◦ C; amount citrate·Mg–Al LDH. Initial uranium concentration 200 mg L−1 ; pH 6.0; temperature
of magnetic citrate·Mg–Al LDH 0.05 g. 25 ◦ C; amount of magnetic citrate·Mg–Al LDH 0.005–0.15 g.

formation of a citrate–metal complexes between uranium(VI) and It can provide more sorption sites to adsorb uranium(VI) ions and
the intercalated citrate. thereby result in the increasing of uranium(VI) ion sorption. From
We have investigated the parameters which affect the ura- the economical point of view, 0.05 g adsorbent dose is selected as
nium(VI) adsorption properties of magnetic citrate·Mg–Al LDH. optimum parameter.

3.2.1. Effect of solution pH


3.2.3. Effect of contact time and adsorption dynamics
The solution pH is one of the most important parameters
Fig. 7 shows the effect of the contact time on uranium(VI)
affecting the adsorption process. In this study, the adsorption
adsorption onto magnetic citrate·Mg–Al LDH composite. The
experiments have been conducted in the initial pH range of
adsorption amount is observed to increase rapidly during start-
2.0–12.0, and the results are shown in Fig. 5. It can be seen from
ing stage of adsorption, then followed by a slowly increase rate
Fig. 5 that the uranium(VI) adsorption capacity increases steeply
and tended to equilibrium at 4 h. The fast uranium(VI) removal
with the increase of pH value, and reaches a maximum when the
rate in the beginning is attributed to the rapid diffusion of ura-
pH value is 6.0. Then, the adsorption capacity turns to decline with
nium(VI) from the solution to the external surfaces of magnetic
the increased pH value beyond 6.0. It may be attributed to the
citrate·Mg–Al LDH. In the slow adsorption process, uranium(VI)
surface charge of magnetic citrate·Mg–Al LDH which could be mod-
ions are presumably adsorbed by chelation with the ligands in the
ified by changing the pH of the solution. At low pH, uncomplexed
interlayer or the ion-exchange in the inner surface of magnetic
UO2 2+ is dominant and the binding sites of magnetic citrate·Mg–Al
citrate·Mg–Al LDH [29]. Moreover, the initial rapid adsorption may
LDH may become positively charged due to the protonation reac-
be because of a large number of available sites at the initial stage.
tion, displaying the cationic character and leading to the low
The more targets of uranium(VI) can provide higher driving force to
adsorption ability of magnetic citrate·Mg–Al LDH for uranium(VI).
facilitate the ions diffusion from the solution to active sites assem-
As pH increases, the functional groups of magnetic citrate·Mg–Al
bly [30,31]. As time proceeds, the concentration gradients became
LDH are progressively deprotonated, forming negative charge. The
reduced owing to the accumulation of uranium(VI) adsorbed on the
attractive forces between the anionic surface sites and cationic
surface sites, leading to the decrease in adsorption rate at the later
uranyl ions easily result in the formation of metal–ligand mag-
stages. Here, to further study the adsorption process, the adsorption
netic composite complexes. However, with further increase of the
pH value, the adsorption capacity decreases. It can be explained
that with the increasing of the concentration of hydroxyl, dis-
180
solved carbonate and bicarbonate anions, uranium(VI) occurs in
the solution in the form of hydroxyl and carbonate-complex ions
175
(UO2 (CO3 )3 4− , (UO2 )2 CO3 (OH)3 − , UO2 (OH)3 − etc.) that are highly
negative charged, leading to low adsorption capacity [28]. On basis 170
Qe (mg/g)

of above results, the optimum pH value is to be 6.0.


165
3.2.2. Effect of adsorbent dose
The effect of the adsorbent dose on the adsorption of uranium 160
was studied. The volume of solution (50 mL) and the concen-
tration of uranium (200 mg L−1 ) were kept constant while the 155
amount of magnetic citrate·Mg–Al LDH varied from 0.005 to 1.5 g.
150
It can be seen from Fig. 6 that the adsorption capacity of ura-
nium(VI) increases rapidly with the increase in adsorbent dose 0 100 200 300 400 500
at m < 0.05 g, and then increases unobviously at m > 0.05 g. The
increase in adsorption with an increase in amount of adsorbent can
Time (min)
be attributed to increased surface area and the availability of more
Fig. 7. Effect of contact time on uranium(VI) adsorption. Initial uranium concentra-
adsorption sites. With increasing magnetic citrate·Mg–Al LDH con- tion 200 mg L−1 ; pH 6.0; temperature 25 ◦ C; amount of magnetic citrate·Mg–Al LDH
tent, the available sites on magnetic citrate·Mg–Al LDH increase. 0.05 g.
224 X. Zhang et al. / Colloids and Surfaces A: Physicochem. Eng. Aspects 414 (2012) 220–227

Table 1
Pseudo-first and pseudo-second-order constants and values of R2 for magnetic citrate·Mg–Al LDH.
exp
Kinetic model T (◦ C) C0 (mg L−1 ) Qe (mg g−1 ) Qecal (mg g−1 ) k1ads (min−1 )/k2ads (mg−1 min−1 ) R2
−2
Pseudo-first order 25 200 180.00 23.05 2.23 × 10 0.94488
Pseudo-second order 25 200 180.00 180.50 3.02 × 10−3 0.99998

data were treated according to kinetic models of pseudo-first-order


and pseudo-second-order [32,33]. The pseudo-first-order kinetic 200
equation is given as:
dQe 160
= k1ads (Qe − Qt ) (2)
dt

Qe (mg/g)
Integrating Eq. (2) and applying the initial conditions, we have 120
25oC
ln(Qe − Qt ) = ln(Qe ) − k1ads t (3) 35oC
80
where k1ads is the rate constant of pseudo-first-order adsorption, Qe 45oC
and Qt (mg g−1 ) refer to the amount of uranium(VI) ions adsorbed
40 55oC
at equilibrium and at time (t), respectively.
The pseudo-second order model considers the rate-limiting step
as the formation chemisorptive bond involving sharing or exchange 0
of electrons between adsorbate and the adsorbent. The model can 0 20 40 60 80
be represented by the following equation: Ce (mg/L)
dQt
= k2ads (Qe − Qt )2 (4) Fig. 9. Adsorption isotherm of magnetic citrate·Mg–Al LDH for uranium(VI)
dt at different temperatures. pH 6.0; temperature 25–55 ◦ C; amount of magnetic
citrate·Mg–Al LDH 0.05 g.
Integrating Eq. (4) and applying the initial conditions, we have
t 1 t
= + (5)
Qt k2ads Qe 2 Qe
where k2ads is the rate constant of pseudo-second-order adsorption.
The values of Qe , k1ads and k2ads are calculated from the intercept
1.5 A and slope values of the plot (Fig. 8A and B) corresponding to Eqs. (3)
and (5), which are given in Table 1. The comparison between the
1.0 experimental adsorption capacity (Qexp ) values and the calculated
adsorption capacity (Qcal ) values shows that pseudo-second-order
adsorption kinetic is very close to the experimental data. More-
ln (Qe-Qt)

0.5
over, the correlation coefficient value for pseudo-second-order
model is much higher than that of pseudo first-order model, sug-
0.0
gesting that the adsorbent systems can be well-described by the
pseudo-second-order kinetic model. As the pseudo-second-order
-0.5 y=-0.00967x+1.36276 model is based on the assumption that chemical sorption is the
2
R =0.94488 rate-determining step [34], the above results obtained consistently
-1.0 suggest that the rate-determining step may be chemical adsorp-
tion and the adsorption behavior may involve the valency forces
0 50 100 150 200 250 through sharing electrons between the uranium(VI) ions and adsor-
t (h) bents [35–37].
3.0
B
2.5
3.4
2.0
t/Qt (hg/mg)

3.2 y=-3408.78764x+13.66084
2
R =0.98523
1.5
3.0

1.0 2.8
ln KD

0.5 y=0.00554x+0.01018 2.6


2
R =0.99998
0.0 2.4

0 100 200 300 400 500 2.2


t (h) 3.0x10
-3
3.1x10
-3 -3
3.2x10 3.3x10
-3
3.4x10
-3

Fig. 8. Pseudo-first-order (A), pseudo-second-order (B) plot for the removal of ura- 1/T (1/K)
nium(VI) by magnetic citrate·Mg–Al LDH. pH 6.0; temperature 25 ◦ C; amount of
magnetic citrate·Mg–Al LDH 0.05 g. Fig. 10. Relationship curve between ln KD and 1/T.
X. Zhang et al. / Colloids and Surfaces A: Physicochem. Eng. Aspects 414 (2012) 220–227 225

0.20
A
0.18

Ce/Qe (g/L)
0.16

0.14

y=-0.00474x+0.20254
0.12
2
R =0.88275

0.10
2 4 6 8 10 12 14 16 18 20 22

Ce (mg/L)
5.5
B
y=1.32936x+1.17667
5.0
2
R =0.99737
4.5
lnQe

4.0

3.5

3.0

1.2 1.4 1.6 1.8 2.0 2.2 2.4 2.6 2.8 3.0 3.2

lnCe
Fig. 11. Langmuir (A) and Freundlich (B) plot for the removal of uranium(VI) by magnetic citrate·Mg–Al LDH. pH 6.0; temperature 25 ◦ C; amount of magnetic citrate·Mg–Al
LDH 0.05 g.

3.2.4. Effect of temperature and adsorption thermodynamics where KD is the distribution coefficient (mL g−1 ), H◦ is standard
To investigate the thermodynamics of adsorption, temperature enthalpy(kJ mol−1 ), S◦ is standard entropy (J mol−1 K−1 ), T is the
experiments were carried out at 25, 35, 45 and 55 ◦ C at pH 6.0 absolute temperature (K), R is the gas constant (8.314 J mol−1 K−1 )
(Fig. 9). It can be observed that the adsorption capacity of ura- and G◦ is the standard Gibbs free energy. The values of H◦ and
nium(VI) ion increases with increasing temperature, indicating that S◦ are calculated from the slope and intercept of the linear regres-
the process is endothermic in nature. The Gibbs energy (G◦ ), sion of ln KD versus 1/T (Fig. 10). The data of G◦ , H◦ and S◦ are
enthalpy (H◦ ), and entropy (S◦ ) are calculated from the follow- together summarized in Table 2.
ing equation: The positive value of H◦ indicates the endothermic nature
o o of the adsorption of uranium(VI) on magnetic citrate·Mg–Al LDH.
H S Hence, the uranium(VI) ions uptake increases with the increase
ln KD = − + (6)
RT R in temperature. The positive value of S◦ suggests the increased
o
G = H − TS
o o
(7) randomness at the solid/liquid interface during the adsorption of

Table 2
Thermodynamics parameters for uranium adsorption on magnetic citrate·Mg–Al LDH.

C0 (mg L−1 ) H◦ (kJ mol−1 ) S◦ (J mol−1 K−1 ) G◦ (kJ mol−1 )

25 ◦ C 35 ◦ C 45 ◦ C 55 ◦ C

200 28.3407 113.5762 −5.5220 −6.6578 −7.8333 −8.9293


226 X. Zhang et al. / Colloids and Surfaces A: Physicochem. Eng. Aspects 414 (2012) 220–227

Table 3 citrate acid anions were likely adsorbed onto the surfaces as well.
Isotherm constants and values of R2 for magnetic citrate·Mg–Al LDH.
The magnetic citrate·Mg–Al LDH is superior to the magnetic Mg–Al
Parameter Value R2 LDH in the uptake of uranium(VI) from solution at a constant
Langmuir isotherm Qm (mg g ) −1
210.97 0.88275 pH of 6.0. This is mainly attributable to the formation of chelate
b (L mg−1 ) 0.0234 complex between uranium(VI) and the intercalated or adsorbed
Freundlich isotherm K (L g−1 ) 3.2436 0.99737 C6 H5 O7 3− . Thus, the citrate in the magnetic Mg–Al LDH has contri-
n 0.7522 bution to the uranium(VI) uptake from solution. The as-prepared
magnetic citrate·Mg–Al LDH composite shows an excellent abil-
uranium(VI) on magnetic citrate·Mg–Al LDH. The Gibbs free energy ity to remove uranium from aqueous solution and the maximum
is found to be decreasing with rise in temperature, which indicates uranium(VI) sorption capacity is about 180.00 mg g−1 at 25 ◦ C. In
the feasibility and spontaneity of the adsorption process. addition, the magnetic citrate·Mg–Al LDH is easily separated by
the external magnetic field; by this way, the problem of phase
3.2.5. Adsorption isotherms of uranium separation using the traditional adsorbents can be resolved. The
The equilibrium data were analyzed using the Langmuir and equilibrium isotherm data are fitted well by Freundlich model and
Freundlich equilibrium models in order to obtain the best fitting pseudo-second-order equation. The thermodynamic parameters
isotherm. The Langmuir equation has been used extensively for G◦ , H◦ and S◦ values of uranium(VI) adsorption onto magnetic
dilute solutions in the following form [38,39] citrate·Mg–Al LDH show the process is endothermic and sponta-
neous in nature. The results of the present investigation illustrate
Ce 1 Ce that magnetic citrate·Mg–Al LDH could be a perfect candidate as an
= + (8)
Qe bQm Qm adsorbent to remove the toxic and radioactive uranium(VI) from
where Ce (mg L−1 ) is the equilibrium concentration of uranium(VI) solution.
ions, Qe (mg g−1 ) is the adsorbed value of uranium(VI) ions at equi-
librium concentration, Qm (mg g−1 ) is the saturated adsorption
amount of uranium(VI), b is Langmuir constant. According to Eq. Acknowledgments
(8), a line is obtained and it is presented in Fig. 11A. The values of
Qm and b can be obtained by the slope and intercept of the line and This work was supported by the Fundamental Research
given in Table 3. Funds of the Central University (HEUCFZ1107), Science and
The essential features of the Langmuir isotherm model can be Technology Planning Project from Education Department of Hei-
expressed in terms of a dimensionless constant separation factor longjiang Province (11553044), High Education Doctoral Fund
or equilibrium parameter RL [40], which is defined by (160100110010), Special Innovation Talents of Harbin Science and
Technology (2010RFXXG007), the foundation of Harbin Engineer-
1
RL = (9) ing University (No. HEUFT07053).
1 + bC0
where C0 is the initial metal ion concentration. The value of sepa-
ration factor constant (RL ) that gives indication for the possibility References
of the adsorption process to proceed: RL > 1, unfavorable; RL = 1,
favorable (linear); 0 < RL < 1, favorable; RL = 0, irreversible; RL < 0, [1] A. Mellah, S. Chegrouche, M. Barkat, The precipitation of ammonium uranyl
unfavorable. carbonate (AUC): thermodynamic and kinetic investigations, Hydrometallurgy
85 (2007) 163–171.
In this study RL value as 0.1761 indicates that magnetic [2] A.T. Kuhu, Electrochemistry of Cleaner Environments, Plenum Press, New York,
citrate·Mg–Al LDH is a suitable adsorbent for adsorption of ura- 1972.
nium(VI) from aqueous solutions. [3] M. Martins, M.L. Faleiro, S. Chaves, R. Tenreiro, E. Santos, M.C. Costa, Anaero-
bic bio-removal of uranium (VI) and chromium (VI): comparison of microbial
The results of uranium(VI) adsorption onto magnetic community structure, J. Hazard. Mater. 176 (2010) 1065–1072.
citrate·Mg–Al LDH have also been analyzed by using the Freund- [4] H. Singh, S.L. Mishra, R. Vijayalakshmi, Uranium recovery from phosphoric acid
lich model to evaluate parameters associated with adsorption by solvent extraction using a synergistic mixture of dinonyl phenyl phosphoric
acid and tri-n-butyl phosphate, Hydrometallurgy 73 (2004) 63–70.
behavior. The Freundlich equation is [41] [5] M.J. Comarmond, T.E. Payne, J.J. Harrison, S. Thiruvoth, H.K. Wong, R.D. Augh-
1/n terson, G.R. Lumpkin, K. Müller, H. Foerstendorf, Uranium sorption on various
Qe = KCe (10) forms of titanium dioxide influence of surface area, surface charge, impurities,
Environ. Sci. Technol. 45 (2011) 5536–5542.
Eq. (10) can be rearranged to linear from [6] N. Pekel, O. Guven, Separation of uranyl ions with amidoximated
poly(acrylonitrile/N-vinylimidazole) complexing sorbents, Colloid Surf., A 212
1
ln Qe = ln K + ln Ce (11) (2003) 155–161.
n [7] Y.J. Jung, S. Kim, S. Park, J.M. Kim, Application of polymer-modified nanoporous
silica to adsorbents of uranyl ions, Colloid Surf., A 313–314 (2008) 162–166.
where K and n are the Freundlich constants, which represent sorp- [8] K. Stamberg, K.A. Venkatesan, P.R. Vasudeva Rao, Surface complexation mod-
tion capacity and sorption intensity, respectively. They can be eling of uranyl ion sorption on mesoporous silica, Colloid Surf., A 221 (2003)
evaluated from the intercept and slope of the linear plot of ln Qe 149–162.
[9] M. Rafatullah, O. Sulaiman, R. Hashim, A. Ahmad, Adsorption of methylene blue
versus ln Ce (Fig. 11B).
on low-cost adsorbents: a review, J. Hazard. Mater. 177 (2010) 70–80.
The adsorption constants of uranium(VI) ions onto magnetic [10] A. Bhatnagar, E. Kumar, M. Sillanp, Fluoride removal from water by
citrate·Mg–Al LDH are calculated according to Langmuir and adsorption—a review, Chem. Eng. J. 171 (2011) 811–840.
[11] S. Chattopadhyay, S.S. Das, A simple and rapid technique for radiochemical
Freundlich adsorption models and the results are listed in Table 3.
separation of iodine radionuclides from irradiated tellurium using an activated
Based on correlation coefficient values (>0.99), it is found that the charcoal column, Appl. Radiat. Isotopes 67 (2009) 1748–1750.
adsorption onto magnetic citrate·Mg–Al LDH can be best described [12] D. James, G. Venkateswaran, T.P. Rao, Removal of uranium from mining indus-
by the Freundlich model. try feed simulant solutions using trapped amidoxime functionality within a
mesoporous imprinted polymer material, Microporous Mesoporous Mater. 119
(2009) 165–170.
4. Conclusions [13] D.M. Sherm, C.L. Peacock, C.G. Hubbar, Surface complexation of uranium(VI) on
goethite (␣-FeOOH), Geochim. Cosmochim. Acta 72 (2008) 298–310.
[14] Y.F. Zhao, M. Wei, J. Lu, Z.L. Wang, X. Duan, Biotemplated hierarchical
Magnetic citrate·Mg–Al LDH, which had C6 H5 O7 3− intercalated nanostructure of layered double hydroxides with improved photocatalysis per-
in the interlayer, was prepared via ion-exchange technique. Some formance, ACS Nano 12 (2009) 4009–4016.
X. Zhang et al. / Colloids and Surfaces A: Physicochem. Eng. Aspects 414 (2012) 220–227 227

[15] F.Z. Zhang, X.F. Zhao, C.H. Feng, B. Li, T. Chen, W. Lu, X.D. Lei, S.L. Xu, Crystal- [28] T.K. Tokunaga, Y. Kim, J.M. Wan, L. Yang, Aqueous uranium(VI) concentra-
face-selective supporting of gold nanoparticles on layered double hydroxide as tions controlled by calcium uranyl vanadate precipitates, Environ. Sci. Technol.
efficient catalyst for epoxidation of styrene, ACS Catal. 1 (2011) 232–237. (2012), http://dx.doi.org/10.1021/es300925u.
[16] L.E. Gaini, M. Lakraimi, E. Sebbar, A. Meghea, M. Bakasse, Removal of indigo [29] M.H. Al-Qunaibit, W.K. Mekhemer, A.A. Zaghloul, The sorption of Cu(II) ions on
carmine dye from water to Mg–Al–CO3 -calcined layered double hydroxides, J. bentonite–a kinetic study, J. Colloid Interface Sci. 283 (2005) 316–321.
Hazard. Mater. 161 (2009) 627–632. [30] H.Y. Xiao, Z.H. Ai, L.Z. Zhang, Nonaqueous sol-gel synthesized hierarchical CeO2
[17] S. Martinez-Gallegos, H. Pfeiffer, E. Lima, M. Espinosa, P. Bosch, S. Bulbu- nanocrystal microspheres as novel adsorbents for wastewater treatment, J.
lian, Cr(VI) immobilization in mixed (Mg,Al) oxides, Microporous Mesoporous Phys. Chem. C 113 (2009) 16625–16630.
Mater. 94 (2006) 234–242. [31] Z.H. Ai, Y. Cheng, L.Z. Zhang, J.R. Qiu, Efficient removal of Cr(VI) from aqueous
[18] S.A. Kulyukhin, E.P. Krasavina, I.V. Gredina, L.V. Mizina, Sorption of U(VI) from solution with Fe@Fe2 O3 core-shell nanowires, Environ. Sci. Technol. 42 (2008)
aqueous solutions on layered double hydroxides of Mg, Al, and Nd, Radiochem- 6955–6960.
istry 52 (2010) 653–661. [32] Y.S. Ho, G. McKay, The kinetics of sorption of divalent metal ions onto sphagnum
[19] S. Vijaikumar, A. Dhakshinamoorthy, K. Pitchumani, L-Proline anchored hydro- moss peat, Water Res. 34 (2000) 735–742.
talcite clays: an efficient catalyst for asymmetric Michael addition, Appl. Catal., [33] M. Yurdakoc, Y. Scki, S.K. Yuedakoc, Kinetic and thermodynamic studies of
A 340 (2008) 25–32. boron removal by Siral 5, Siral 40, and Srial 80, J. Colloid Interface Sci. 286
[20] L. Ingram, H.F.W. Taylor, The crystal structures of sjogrenite and pyroaurite, (2005) 440–446.
Miner. Mag. 36 (1967) 465–479. [34] C. Chen, J.L. Wang, Removal of Pb2+ , Ag+ , Cs+ and Sr2+ from aqueous solution by
[21] M.R. Perez, I. Pavlovic, C. Barriga, J. Cornejo, M.C. Hermosin, M.A. Ulibarri, brewery’s waste biomass, J. Hazard. Mater. 151 (2008) 65–70.
Uptake of Cu2+ ,Cd2+ and Pb2+ on Zn–Al layered double hydroxide intercalated [35] Y.S. Ho, G. McKay, Sorption of dye from aqueous solution by peat, Chem. Eng.
with EDTA, Appl. Clay Sci. 32 (2006) 245–251. J. 70 (1998) 115–124.
[22] R. Rojas, M.R. Perez, E.M. Erro, P.I. Ortiz, M.A. Ulibarri, C.E. Giacomelli, EDTA [36] A.K. Bhattacharyal, T.K. Naiya, S.N. Mondal, S.K. Das, Adsorption, kinetics and
modified LDHs as Cu2+ scavengers: removal kinetics and sorbent stability, J. equilibrium studies on removal of Cr(VI) from aqueous solutions using different
Colloid Interface Sci. 331 (2009) 425–431. low-cost adsorbents, Chem. Eng. J. 137 (2008) 529–541.
[23] N.H. Gutmann, L. Spiccia, T.W. Turney, Complexation of Cu(II) and Ni(II) by [37] L.M. Zhou, J.Y. Jin, Z.R. Liu, X.Z. Liang, C. Shang, Adsorption of acid dyes
nitrilotriacetate intercalated in Zn–Cr layered double hydroxides, J. Mater. from aqueous solutions by the ethylenediamine-modified magnetic chitosan
Chem. 10 (2000) 1219–1224. nanoparticles, J. Hazard. Mater. 185 (2011) 1045–1052.
[24] J.W. Qian, D.P. Li, G.Q. Zhan, L. Zhang, W.T. Su, P. Gao, Simultaneous biodegra- [38] N. Sharma, K. Kaur, S. Kaur, Kinetic and equilibrium studies on the removal of
dation of Ni–citrate complexes and removal of nickel from solutions by Cd2+ ions from water using polyacrylamide grafted rice (Oryza sativa) huskand
Pseudomonas alcaliphila, Bioresour. Technol. 116 (2012) 66–73. (Tectona grandis) saw dust, J. Hazard. Mater. 163 (2009) 1338–1344.
[25] Y. Hu, G.Q. Shen, H.L. Zhu, G.X. Jiang, A class-specific enzyme-linked [39] S. Chegrouche, A. Mellah, S. Telmoune, Removal of lanthanum from aqueous
immunosorbent assay based on magnetic particles for multiresidue solutions by natural bentonite, Water Res. 31 (1997) 1733–1737.
organophosphorus pesticides, J. Agric. Food Chem. 58 (2010) 2801–2806. [40] A. Bhatnagar, A.K. Jain, A comparative adsorption study with different industrial
[26] Y. Zhao, F. Li, R. Zhang, D.G. Evans, X. Duan, Preparation of layered double- wastes as adsorbents for the removal of cationic dyes from water, J. Colloid
hydroxide nanomaterials with a uniform crystallite size using a new method Interface Sci. 28 (2005) 49–55.
involving separate nucleation and aging steps, Chem. Mater. 14 (2002) [41] X.L. Tan, X.K. Wang, M. Fang, C.L. Chen, Sorption and desorption of Th(IV) on
4286–4291. nanoparticles of anatase studied by batch and spectroscopy methods, Colloid
[27] J. Zhang, F.Z. Zhang, L.L. Ren, D.G. Evans, X. Duan, Synthesis of layered double Surf., A 296 (2007) 109–116.
hydroxide anionic clays intercalated by carboxylate anions, Mater. Chem. Phys.
85 (2004) 207–214.

Вам также может понравиться